mer, fac, and Bidentate Coordination of an Alkyl-POP Ligand in the

Dec 13, 2016 - Nonclassical and classical osmium polyhydrides containing the diphosphine 9,9-dimethyl-4,5-bis(diisopropylphosphino)xanthene (xant(PiPr...
0 downloads 9 Views 1MB Size
Article pubs.acs.org/IC

mer, fac, and Bidentate Coordination of an Alkyl-POP Ligand in the Chemistry of Nonclassical Osmium Hydrides Miguel A. Esteruelas,* Cristina García-Yebra, Jaime Martín, and Enrique Oñate Departamento de Química Inorgánica, Instituto de Síntesis Química y Catálisis Homogénea (ISQCH), Centro de Innovación en Química Avanzada (ORFEO-CINQA), Universidad de Zaragoza-CSIC, 50009 Zaragoza, Spain S Supporting Information *

ABSTRACT: Nonclassical and classical osmium polyhydrides containing the diphosphine 9,9-dimethyl-4,5-bis(diisopropylphosphino)xanthene (xant(PiPr2)2), coordinated in κ3-mer, κ3-fac, and κ2-P,P fashions, have been isolated during the cyclic formation of H2 by means of the sequential addition of H+ and H− or H− and H+ to the classical trihydride OsH3Cl{xant(PiPr2)2} (1). This complex adds H+ to form the compressed dihydride dihydrogen complex [OsCl(H···H)(η2H2){xant(PiPr2)2}]+ (2). Under argon, cation 2 loses H2 and the resulting unsaturated fragment dimerizes to give [(Os(H···H){xant(PiPr2)2})2(μ-Cl)2]2+ (3). During the transformation the phosphine changes its coordination mode from mer to fac. The benzofuran counterpart of 1, OsH3Cl{dbf(PiPr2)2} (4; dbf(PiPr2)2 = 4,6-bis(diisopropylphosphino)dibenzofuran), also adds H+ to afford the benzofuran counterpart of 2, [OsCl(H··· H)(η2-H2){xant(PiPr2)2}]+ (5), which in contrast to the latter is stable and does not dimerize. Acetonitrile breaks the chloride bridge of 3 to form the dihydrogen [OsCl(η2-H2)(CH3CN){xant(PiPr2)2}]+ (6), regenerating the mer coordination of the diphosphine. The hydride ion also breaks the chloride bridge of 3. The addition of KH to 3 leads to 1, closing a cycle for the formation of H2. Complex 1 reacts with a second hydride ion to give OsH4{xant(PiPr2)2} (7) as consequence of the displacement of the chloride. Similarly to the latter, the oxygen atom of the mer-coordinated diphosphine of 7 has a tendency to be displaced by the hydride ion. Thus, the addition of KH to 7 yields [OsH5{xant(PiPr2)2}]− (8), containing a κ2-P,Pdiphosphine. Complex 8 is easily protonated to afford OsH6{xant(PiPr2)2} (9), which releases H2 to regenerate 7, closing a second cycle for the formation of molecular hydrogen.



diphosphines that could act as κ3-fac and bidentate groups, in addition to κ3-mer, flexible POP ethers have been also employed.5 Thus, Weller and co-workers have demonstrated that 4,5-bis(diphenylphosphino)-9,9-dimethylxanthene (xantphos) has the ability of changing its coordination fashion in several relevant rhodium- and iridium-mediated processes,6 such as the hydroacylation7 and carbothiolation8 of alkenes and alkynes and the dehydrocoupling of amine−boranes.9 Transition-metal polyhydrides are compounds having enough hydrogen atoms bonded to the metal center of a LnM fragment to form at least two different types of ligands.10 These ligands are generally classified into four types depending upon the separation between the coordinated hydrogens: Kubas type dihydrogens (0.8−1.0 Å), elongated dihydrogens (1.0−1.3 Å), compressed dihydrides (1.3−1.6 Å), and classical hydrides (>1.6 Å).10,11 Polyhydrides of the platinum-group metals offer new exciting conceptual challenges and the possibility of interacting with different fields,10 including the conversion and storage of regenerative energy.12 In this respect, osmium polyhydrides are of particular interest. In addition to promoting the dehydrocoupling of amine−boranes,13 their

INTRODUCTION The chemical behavior of a transition-metal complex is determined by the central ion and by the ligands forming its coordination sphere. The groups surrounding the core govern the available electron density of the metal ion and the accessible space for performing the reactions. In some cases, the ligands also cooperate with the metal by means of a direct participation in the chemical transformations.1 There are other ligands that change their properties during the reactions. This group includes ligands that modify their electronic donor ability while they are adapted to the reaction medium by means of reversible transformations2 and ligands that change their coordination mode to meet with the requirements of each stage of a multistep process. Hemilabile groups are the simplest class of ligands with the latter behavior. The reversible coordination−decoordination of the hemilabile donor atom allows the stabilization of highly reactive metal centers, which has a significant influence on the catalytic behavior of the complexes.3 Tridentate anionic PCPdiphosphines are situated in the opposite position. Although their mer coordination allows them to stabilize complexes with unusual reactivity,4 the stability and rigidity of the pincer diminishes the adaptability of these ligands to the requirement of the sequential processes. In the search for more versatile © XXXX American Chemical Society

Received: November 23, 2016

A

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Scheme 1. Formation of H2 by Sequential Addition of H+ and H− or H− and H+ to OsH3Cl{xant(PiPr2)2} (1) in Two Related Cycles

dichloromethane-d2 leads to the cation [OsCl(H···H)(η2H2){xant(PiPr2)2}]+ (2). The presence of four hydrogen atoms bonded to the metal center is supported by the 1H NMR spectrum of the resulting solution, which contains a broad signal at −10.71 ppm with an integrated intensity of 4. According to the existence of nonclassical interactions between the coordinated hydrogen atoms, this resonance exhibits a 400 MHz T1(min) value of 40 ± 2 ms at 203 K. The 31P{1H} NMR spectrum shows a singlet at 62.6 ppm. DFT calculations (energies calculated at the B3LYP(GD3)//6-31G(d,p)/SDD level with the Gaussian 09 program23) reveal that there are two compressed dihydride dihydrogen structures differing by 3.0 kcal mol−1 (ΔG, 1 atm, 298.15 K): 2a and 2b. Figure 1 shows a

ability to activate C−H, N−H, and C−N bonds of a wide range of organic molecules, including 2-azetidinones14 and nucleobases,15 allows them to interact with organic synthetic chemistry,16 drug design,14a and materials science.17 Although the chemistry of the osmium polyhydrides is rich, the complexes of this type with pincer ligands are a noticeable exception,10 being reduced to a few complexes with PNP,18 P(olefin)P,19 and P(C(sp3))P20 ligands. Recently, we have reported that the reactions of complexes OsCl2{xant(PiPr2)2}(κ-S-DMSO) (xant(P i Pr 2 ) 2 = 9,9-dimethyl-4,5-bis(diisopropylphosphino)xanthene) and OsCl2{dbf(PiPr2)2}(κS-DMSO) (dbf(PiPr2)2 = 4,6-bis(diisopropylphosphino)dibenzofuran) with molecular hydrogen, in the presence of Et3N, lead to the classical trihydride derivatives OsH3Cl{xant(PiPr2)2}21 and OsH3Cl{dbf(PiPr2)2}16d containing a pincer POP ligand. Our interest in determining the influence of the different coordination modes of flexible pincer ligands on the nature of the H-donor ligands of osmium polyhydrides prompted us to study the protonation of the hydride ion, kinetically controlled by OsH3Cl{xant(PiPr2)2}. The kinetically controlled formation of molecular hydrogen by the exothermic neutralization of the hydride ion is a process of great interest in connection with the storage of hydrogen in solid materials, which would become a safe and efficient way to store energy for both stationary and mobile applications.22 This paper shows that the ligand 9,9-dimethyl-4,5-bis(diisopropylphosphino)xanthene changes its coordination mode to stabilize Kubas type dihydrogens, elongated dihydrogens, compressed dihydrides, and classical hydrides, which act as intermediates to generate molecular hydrogen from H+ and H−, in the presence of osmium polyhydrides. In addition, it proves that the mer coordination favors nonclassical interactions between the coordinated hydrogen atoms, with regard to the fac and bidentate coordination modes.



Figure 1. DFT optimized structures of 2a and 2b. Hydrogen atoms (except hydrides) are omitted for clarity. Selected bond lengths (Å) and angles (deg): for 2a, Os−O = 2.307, Os−P(1) = 2.381, Os−P(2) = 2.384, Os−Cl = 2.438, H(1)−H(2) = 1.427, H(3)−H(4) = 1.029, O−Os−P(1) = 80.0, O−Os−P(2) = 80.2, P(1)−Os−P(2) = 160.2; for 2b, Os−O = 2.224, Os−P(1) = 2.390, Os−P(2) = 2.391, Os−Cl = 2.474, H(1)−H(2) = 1.091, H(3)−H(4) = 1.358, O−Os−P(1) = 82.6, O−Os−P(2) = 82.4, P(1)−Os−P(2) = 165.5.

RESULTS AND DISCUSSION

view of these DFT-optimized structures. In both cases, the diphosphine is mer-coordinated with respective P(1)−Os− P(2), O−Os−P(1), and O−Os−P(2) angles of 160.2, 80.0, and 80.2° for 2a and 165.5, 82.6, and 82.4° for 2b. The most stable structure 2a shows the dihydrogen ligand (1.029 Å) at the plane perpendicular to the P−Os−P direction, disposed trans to the chloride, whereas the compressed dihydride (1.427 Å) lies trans to the oxygen atom of the diphosphine, almost parallel to the P−Os−P direction. In 2b, the dihydrogen (1.091 Å) and

Scheme 1 summarizes the procedures used to generate molecular hydrogen starting from the classical trihydride OsH3Cl{xant(PiPr2)2} (1). Route a involves the sequential addition of H+ and H−, whereas route b involves the opposite addition. Complex 1 adds a proton, in agreement with the Lewis base character of the saturated polyhydrides. Under a hydrogen atmosphere, the addition of 1.0 equiv of HBF4·OEt2 to 1 in B

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry the compressed dihydride (1.358 Å) ligands exchange their positions. The lability of the dihydrogen ligand in 2 suggests that it is of the Kubas type even though the separation between the two hydrogen atoms lies at the border between a Kubas-type dihydrogen and an elongated dihydrogen. Thus, complex 2 can be kept in dichloromethane, under a hydrogen atmosphere, at room temperature, for a few hours. However, under argon, it loses the dihydrogen ligand to afford the dimer [(Os(H··· H){xant(PiPr2)2})2(μ-Cl)2]2+ (3), which was isolated as its yellow BF4− salt in 95% yield and characterized by X-ray diffraction analysis.24 The structure (Figure 2) shows a change

Complex 5 was characterized by X-ray diffraction analysis. Its structure (Figure 3) is in full agreement with the optimized

Figure 3. ORTEP diagram of complex 5 with 50% probability ellipsoids. The counteranions, solvent molecules, and hydrogen atoms (except hydrides) are omitted for clarity. Selected bond lengths (Å) and angles (deg): Os−O = 2.092(3), Os−P(1) = 2.400(1), Os−P(2) = 2.403(1), H(3)−H(4) = 1.06(5), H(1)−H(2) = 1.10(7), Os−Cl = 2.415(1), O−Os−(centroid H(1)−H(2)) = 171.37, Cl−Os−(centroid H(3)−H(4)) = 178.09, O−Os−P(1) = 78.66(9), O−Os−P(2) = 78.72(9), P(1)−Os−P(2) = 157.23(4).

Figure 2. ORTEP diagram of complex 3 with 50% probability ellipsoids. The counteranions, solvent molecules, and hydrogen atoms (except hydrides) are omitted for clarity. Selected bond lengths (Å) and angles (deg): Os−O = 2.156(3), Os−P(1) = 2.316(1), Os−P(2) = 2.295(1), H(1)−H(2) = 1.36(6); O−Os−(centroid H(1)−H(2)) = 169.65, O−Os−P(1) = 80.45(9), O−Os−P(2) = 82.32(9), P(1)− Os−P(2) = 107.72(5).

structure of the most stable isomer of 2. As in the latter, the diphosphine is mer-coordinated with O−Os−P(1), O−Os− P(2), and P(1)−Os−P(2) angles of 78.66(9), 78.72(9), and 157.23(4)°, respectively. The dihydrogen ligand (H(3)−H(4) = 1.06(5) Å (X-ray), 0.976 Å (DFT)) lies at the plane perpendicular to the P−Os−P direction, disposed trans to the chloride; the compressed dihydride (1.10(7) Å (X-ray), 1.387 Å (DFT)) is situated trans to the oxygen atom of the diphosphine, almost parallel to the P−Os−P direction. In the 1 H NMR spectrum in dichloromethane-d2, the OsH4 unit displays a broad signal at −10.60 ppm, which exhibits a 400 MHz T1(min) value of 40 ± 2 ms at 203 K, in agreement with 2. The 31P{1H} NMR spectrum contains a singlet at 65.3 ppm. The coordination changes from κ3-mer to κ3-fac of the diphosphine is a requirement for the coordinative saturation of the metal center on the [OsClH2{xant(PiPr2)}]+ fragment resulting from the dissociation of molecular hydrogen from 2, since the steric hindrance experienced by the isopropyl substituents of the mer-coordinated diphosphines of two unsaturated fragments appears to prevent the dimerization. However, the coordination κ3-mer is clearly favored over κ3-fac, when the metal center of a mononuclear complex is saturated. As a proof of concept, it should be mentioned that the chloride bridges of 3 are broken in acetonitrile to form the saturated dihydrogen complex [OsCl(η2-H2)(CH3CN){xant(PiPr2)2}]BF4 (6), which contains a κ3-mer-coordinated diphosphine. This compound was isolated as a white solid, in almost quantitative yield, and characterized by X-ray diffraction analysis. The structure26 (Figure 4) proves the mer coordina-

in the coordination mode of the diphosphine, which now coordinates in a κ3-fac fashion with O−Os−P(1), O−Os−P(2), and P(1)−Os−P(2) angles of 80.45(9), 82.32(9), and 107.72(5)°, respectively. The coordinated hydrogen atoms, separated by 1.36(6) Å, lie trans to the oxygen atoms of the diphosphines, forming a compressed dihydride. The nonclassical nature of the OsH2 units was confirmed by the 1H NMR spectrum in dichloromethane-d2, which shows a triplet (2JH−P = 14.5 Hz) at −9.53 ppm. In agreement with the H−H separation found by X-ray diffraction analysis, this resonance exhibits a 400 MHz T1(min) value of 63 ± 3 ms, at 243 K, and a JH‑D value of 7 Hz, which fit with H−H separations of 1.35 and 1.32 Å, respectively.25 The 31P{1H} NMR spectrum shows a signal at 42.9 ppm. The fac coordination of the diphosphine in 3 is a determining factor for the stabilization of the dimeric structure with regard to 2, which is related to the flexibility of the xanthene group. In fact, the most rigid benzofuran stabilizes the OsH4 species and prevents the formation of a counterpart dimer of 3. Thus, under argon, the addition of 1.0 equiv of HBF4·OEt2 to OsH3Cl{dbf(PiPr2)2} (4) in dichloromethane leads to the compressed dihydride dihydrogen complex [OsCl(H···H)(η2H2){dbf(PiPr2)2}]BF4 (5), which was isolated as a pale yellow solid in 85% yield (eq 1). In contrast to 2, this 4,6bis(diisopropylphosphino)dibenzofuran counterpart is moderately stable under argon, in the solid state, and in dichloromethane and does not release molecular hydrogen to dimerize. C

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry

Figure 4. ORTEP diagram for one of the two molecules in the symmetry unit of complex 6 with 50% probability ellipsoids. The counteranions, solvent molecules, and hydrogen atoms (except hydrides) are omitted for clarity. Selected bond lengths (Å) and angles (deg): Os−O = 2.131(9), Os−P(1) = 2.338(3), Os−Cl = 2.441(4), Os−N = 2.024(16), H(1)−H(2) = 1.02(6), Cl−Os− (centroid H(1)−H(2)) = 174.96, O−Os−P(1) = 83.08(6), P(1)− Os−P(1) = 165.85(13).

Figure 5. ORTEP diagram of complex 8 with 50% probability ellipsoids. The solvent molecules and hydrogen atoms (except hydrides) are omitted for clarity. Selected bond lengths (Å) and angles (deg): Os−O = 3.496(3), Os−P(1) = 2.309(1), Os−P(2) = 2.322(1), Os−K = 3.624(1), P(1)−Os−P(2) = 109.46(3), P(1)−Os− H(5) = 165.6(16).

tion (O−Os−P(1) = 83.08(6), 83.13(7)°; P(1)−Os−-P(1) = 165.85(13), 165.81(14)°) of the diphosphine. The dihydrogen ligand (1.02(6) Å (X-ray), 0.985 Å (DFT)), which is almost parallel to the P−Os−P direction, lies trans to the chloride. In the 1H NMR spectrum, in dichloromethane-d2, the coordinated hydrogen molecule displays a triplet (2JH−P = 6 Hz) at −9.86 ppm. As expected for its Kubas type dihydrogen nature, this resonance exhibits a 300 MHz T1(min) value of 21 ± 1 ms, at 203 K, whereas the H−D coupling constant in the partially deuterated species is 20 Hz. These values allow us to calculate H−H separations of 0.94 and 1.09 Å, respectively,25 which agree well with those obtained by X-ray diffraction analysis. The 31 1 P{ H} NMR spectrum shows a singlet at 27.2 ppm. The hydride ion also breaks the chloride bridges of 3. The addition of KH to the tetrahydrofuran solutions of the dimer regenerates 1, closing a cycle for the formation of molecular hydrogen by reaction of H+ and H− on the osmium center of 1 (route a in Scheme 1). The hydride ion displaces the chloride ligand of 1 to afford the tetrahydride derivative OsH4{xant(PiPr2)2} (7). The classical nature of the coordinated hydrogen atoms of this previously reported compound was confirmed by X-ray diffraction analysis and DFT calculations.21 Similarly to chloride, the oxygen atom of the mer-coordinated diphosphine of the latter has a tendency to be displaced by the hydride ion. The addition of KH to 7 in tetrahydrofuran leads to the anionic pentahydride [OsH5{xant(PiPr2)2}]− (8), containing a κ2-P,Pbidentate diphosphine. In the presence of the crown ether 18crown-6, yellow crystals of the [K(18-crown-6)]+ salt were obtained in 66% yield. The X-ray diffraction structure (Figure 5) proves the bidentate coordination of the diphosphine, which has a P−Os−P bite angle of 109.46(3)°. The coordination polyhedron around the osmium atom can be idealized as a pentagonal bipyramid with the hydride H(5) and the P(1) atom of the diphosphine in axial positions (P(1)−Os−H(5) = 165.6(16)°), whereas the P(2) atom and the remaining hydrides lie in the equatorial plane. The disposition of the phosphorus atoms is certainly enforced by the bidentate character of the diphosphine. In contrast to 8, the related anion [OsH5(PiPr3)2]− contains both monodentate phosphines at the axial positions of the bipyramid.27 The structure also shows

three hydrides pointing toward the potassium cation, which is slightly out of the plane of the crown ether. The ion pair is helped together through electrostatic interactions, enhanced by the ionic nature of cation and anion. In tetrahydrofuran the hydride ligands, as well as the phosphorus atoms of the diphosphine, rapidly exchange their positions even at 183 K. In agreement with this, the 1H NMR spectrum shows a broad signal at −12.46 ppm, whereas the 31P{1H} NMR spectrum contains a singlet at 20.6 ppm. The classical hydride character of the coordinated hydrogen atoms is strongly supported by the 400 MHz T1(min) value of the hydride resonance, 248 ms, at 208 K. Complex 8 is a strong Brønsted base which is even protonated by traces of water. The protonation leads to the hexahydride OsH6{xant(PiPr2)2} (9), also containing a P,Pbidentate diphosphine with the P atoms at a B site of the BAAB trapezoidal planes, which define its dodecaedral structure.21,28 Complex 9 releases molecular hydrogen and coordinates the oxygen atom to regenerate the tetrahydride 7, closing a second cycle for the formation of molecular hydrogen by reaction of H− with H+ on the metal center of the tetrahydride 7 (route b in Scheme 1).



CONCLUDING REMARKS This study reveals that osmium polyhydrides stabilized by the POP-pincer ligand 9,9-dimethyl-4,5-bis(diisopropylphosphine)xanthene sequentially add H+ and H− or H− and H+ to generate molecular hydrogen in a cyclic manner. During the process, the diphosphine changes its coordination mode from κ3-mer to κ3fac and κ2-P,P-bidentate, depending upon the electronic and steric requirements of the involved intermediates, to form both polyhydrides with nonclassical interactions and classical polyhydrides. The nonclassical interactions are favored by the tridentate coordination, in particular by the mer disposition, whereas the classical polyhydrides are usually found when the diphosphine is bidentate. The flexibility of the central xanthene group is a determining factor for the behavior of the diphosphine. D

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



PCH(CH3)2), 18.4 (d, 2JCP = 4.2, PCH(CH3)2). 31P {1H} NMR (202.46 MHz, CD2Cl2, 293 K): δ 42.9 (br, POP). T1(min) (ms, OsH, 400 MHz, CD2Cl2, 243 K): 63 ± 3 (−9.53 ppm). Determination of the JH‑D Value for Complex 3. A solution of DBF4·D2O was prepared by stirring HBF4 (1 mL) and D2O (1 mL) for 1 h. Then, a solution of [OsH3Cl{xant(PiPr2)2}] (1; 45 mg, 0.067 mmol) in CD2Cl2 (0.5 mL) was placed in an NMR tube and treated with a freshly prepared solution of DBF4 in D2O (22 μL; 0.080 mmol). After 24 h, the 1H{31P} NMR spectrum of the reaction mixture showed a triplet (JH‑D = 7.0 Hz), at −9.53 ppm, corresponding to the deuterated isotopomer of complex 3. Synthesis of [OsCl(H···H)(η2-H2){dbf(PiPr2)2}]BF4 (5). HBF4· OEt2 (14 μL; 0.103 mmol) was added to a solution of [OsH3Cl{dbf(PiPr2)2}] (4; 60 mg, 0,095 mmol) in dichloromethane (3 mL). After the mixture was stirred for 15 min, at room temperature, the solvent was evaporated under reduced pressure and cold pentane (1 mL) was added to afford a pale yellow solid which was washed once with cold pentane (3 mL) and dried overnight under vacuum. Yield: 58 mg (85%). In order to confirm the structure of the cation, crystals suitable for X-ray diffraction analysis were obtained by slow diffusion of pentane into a dichloromethane solution of the complex [OsCl(H··· H)(η2-H2){dbf(PiPr2)2}][OTf]. Anal. Calcd for C24H38BClF4OOsP2: C, 40.20; H, 5.34. Found: C, 39.99; H, 5.43. HRMS (electrospray, m/ z): calcd for C24H37OOsP2 [M − HCl]+, 595.1930; found, 595.1935. IR (ATR, cm−1): ν (BF) 1049 (m). 1H NMR (300 MHz, CD2Cl2, 293 K): δ 8.25 (d, 3JHH = 7.7, 2H, CH-arom), 8.00 (m, 2H, CH-arom), 7.77 (dd, 3JHH = 7.7, 3JHH = 7.6, 2H, CH-arom), 3.27 (m, 2H, PCH(CH3)2), 2.89 (m, 2H, PCH(CH3)2), 1.75 (dvt, N = 16.2, 3JHH = 7.3, 6H, PCH(CH3)2), 1.57 (dvt, N = 19.9, 3JHH = 7.5, 6H, PCH(CH3)2), 1.39 (dvt, N = 19.0, 3JHH = 6.9, 6H, PCH(CH3)2), 0.91 (dvt, N = 18.1, 3JHH= 7.0, 6H, PCH(CH3)2), −10.60 (br, 4H, OsH). 13 C{1H}-APT NMR (75.47 MHz, CD2Cl2, 293 K): δ 161.0 (vt, N = 13.1, C-arom), 131.1 (s, CH-arom), 128.7 (vt, N = 5.4, CH-arom), 127.8 (s, CH-arom), 123.8 (vt, N = 6.5, C-arom), 115.0 (vt, N = 37.6, C-arom), 25.8 (vt, N = 30.7, PCH(CH3)2), 25.2 (vt, N = 29.3, PCH(CH3)2), 22.2 (s, PCH(CH3)2), 20.9 (vt, N = 4.5, PCH(CH3)2), 20.2 (s, PCH(CH3)2), 18.0 (s, PCH(CH3)2). 31P{1H} NMR (121.49 MHz, CD2Cl2, 293 K): δ 65.3 (s, POP). T1(min) (ms, OsH, 400 MHz, CD2Cl2, 203 K): 40 ± 2 (−10.60 ppm). Synthesis of [OsCl(η2-H2)(CH3CN){xant(PiPr2)2}][BF4] (6). [(Os(H···H){xant(PiPr2)2})2(μ-Cl)2][BF4]2 (3; 100 mg; 0.066 mmol) was dissolved in CH3CN (3 mL) and stirred for 30 min, after which the solvent was evaporated. The residue was precipitated with cold pentane (2 × 1 mL) to afford a white solid which was dried overnight under reduced pressure. Yield: 93 mg (99%). Colorless crystals suitable for X-ray diffraction were obtained by vapor diffusion of pentane into a CH2Cl2 solution. Anal. Calcd for C29H45BClF4NOOsP2: C, 43.64; H, 5.68; N, 1.76. Found: C, 43.26; H, 5.72; N, 1.79. IR (ATR, cm−1): ν(C−N) 2288 cm−1; ν(BF) 1048 (s). 1H NMR (300 MHz, CD2Cl2, 293 K): δ 7.64 (m, 2H, CH-arom), 7.57 (dd, 3JHH = 7.6, 4JHH = 1.3, 2H, CH-arom), 7.40 (dd, 3JHH = 7.6, 3 JHH = 7.5, 2H, CH-arom), 3.32 (m, 2H, PCH(CH3)2), 2.86 (m, 2H, PCH(CH3)2), 2.77 (s, 3H, CH3CN), 1.65 (dvt, N = 15.1, 3JHH = 7.2, 6H, PCH(CH3)2), 1.61 (s, 3H, C(CH3)2), 1.59 (s, 3H, C(CH3)2), 1.54 (dvt, N = 16.9, 3JHH = 7.5, 6H, PCH(CH3)2), 1.42 (dvt, N = 17.2, 3 JHH = 6.9, 6H, PCH(CH3)2), 0.94 (dvt, N = 16.4, 3JHH = 6.1, 6H, PCH(CH3)2), −9.86 (t, 2JHP = 6.0, 2H, OsH). 13C{1H}-APT NMR (75.47 MHz, CD2Cl2, 293 K): δ 158.2 (vt, N = 11.6, C-arom), 132.4 (vt, N = 6.0, C-arom), 132.1 (s, CH-arom), 131.2 (s, CH-arom), 127.3 (vt, N = 6.3, CH-arom), 124.4 (s, CH3CN), 123.7 (vt, N = 38.4, Carom), 34.2 (s, C(CH3)2) 34.1 (s, C(CH3)2), 33.1 (s, C(CH3)2), 28.0 (vt, N = 28.1, PCH(CH3)2), 27.9 (vt, N = 32.0, PCH(CH3)2), 21.8 (s, PCH(CH3)2), 19.8 (vt, N = 5.9, PCH(CH3)2), 19.8 (s, PCH(CH3)2), 18.7 (s, PCH(CH3)2), 4.2 (s, CH3CN). 31P{1H} NMR (121.49 MHz, CD2Cl2, 293 K): δ 27.2 (s, POP). T1(min) (ms, OsH, 300 MHz, CD2Cl2, 203 K): 21 ± 1 (−10.51 ppm). Determination of the JHD Value for complex 6. DOTf (3 μL; 0.034 mmol) was added to a solution of [OsH3Cl{xant(PiPr2)2}] (1; 23 mg; 0.034 mmol) in CD2Cl2 (0.5 mL). Then, CH3CN (2 μL; 0.038 mmol) was added. The 1H{31P} NMR spectrum of this solution

EXPERIMENTAL SECTION

General Information. All reactions were carried out under argon with rigorous exclusion of air using Schlenk tube or glovebox techniques. Solvents were dried by the usual procedures and distilled under argon prior to use or obtained oxygen- and water-free from an MBraun solvent purification apparatus. The starting materials [OsH3Cl{xant(PiPr2)2}] (1),21 [OsH3Cl{dbf(PiPr2)2}] (4),16d and [OsH4{xant(PiPr2)2}] (7)21 were prepared according to published methods. 1H, 31P{1H}, and 13C{1H} NMR spectra were recorded on a Bruker 300 ARX, Bruker Avance 300 MHz, or Bruker Avance 400 MHz instrument. Chemical shifts (expressed in parts per million) are referenced to residual solvent peaks (1H, 13C{1H}) or external H3PO4 (31P{1H}). Coupling constants, J and N (N = JPH + JP′H for 1H; N = JPC + JP′C for 13C), are given in hertz. Spectral assignments were achieved by 1H−1H COSY, 1H{31P}, 13C APT, 1H−13C HSQC, and 1 H−13C HMBC experiments. Infrared spectra were recorded on a PerkinElmer Spectrum One or PerkinElmer Spectrum 100 FT-IR spectrometer, equipped with an ATR accessory, as neat solids. C, H, and N analyses were carried out in a PerkinElmer 2400 CHNS/O analyzer. High-resolution electrospray mass spectra were acquired using a MicroTOF-Q hybrid quadrupole time-of-flight spectrometer (Bruker Daltonics, Bremen, Germany). Formation of [OsCl(H···H)(η2-H2){xant(PiPr2)2}]BF4 (2). A solution of [OsH3Cl{xant(PiPr2)2}] (1; 50 mg; 0.073 mmol) in dichloromethane-d2 (0.6 mL) was placed in a screw-top NMR tube under an H2 atmosphere, and HBF4·OEt2 (11 μL; 0.081 mmol) was added. NMR spectra of the reaction solution showed immediate and quantitative formation of 2. All attempts to isolate the new species were unsuccessful due to easy loss of hydrogen to form the dinuclear complex 3. HRMS (electrospray, m/z): calcd for C27H44ClOOsP2 [M]+, 673.2158; found, 673.2155. 1H NMR (400 MHz, CD2Cl2, 233 K): δ 7.76 (d, 3JHH = 7.5, 2H, CH-arom), 7.67 (m, 2H, CH-arom), 7.53 (dd, 3 JHH = 7.6, 3JHH = 7.5, 2H, CH-arom), 3.09 (m, 2H, PCH(CH3)2), 2.79 (m, 2H, PCH(CH3)2), 1.81 (s, 3H, C(CH3)2), 1.60 (dvt, N = 16.4, 3JHH = 7.0, 6H, PCH(CH3)2), 1.56 (s, 3H, C(CH3)2), 1.45 (dvt, N = 18.4, 3JHH = 7.6, 6H, PCH(CH3)2), 1.30 (dvt, N = 18.1, 3JHH = 7.1, 6H, PCH(CH3)2), 0.75 (dvt, N = 17.2, 3JHH = 7.1, 6H, PCH(CH3)2), −10.71 (br, 4H, OsH). 13C{1H}-APT NMR (100.62 MHz, CD2Cl2, 233 K): δ 155.4 (vt, N = 12.5, C-arom), 132.4 (vt, N = 5.9, C-arom), 132.0 (s, CH-arom), 131.9 (s, CH-arom), 127.7 (vt, N = 6.2, CH-arom), 119.8 (vt, N = 39.4, C-arom), 35.0 (s, C(CH3)2), 34.0 (s, C(CH3)2), 30.3 (s, C(CH3)2), 26.6 (vt, N = 33.1, PCH(CH3)2), 26.0 (vt, N = 35.4, PCH(CH3)2), 22.1 (s, PCH(CH3)2), 19.8 (s, PCH(CH3)2), 19.2 (s, PCH(CH3)2), 18.3 (s, PCH(CH3)2). 31P{1H} NMR (161.98 MHz, CD2Cl2, 233 K): δ 62.6 (s, POP). T1(min) (ms, OsH, 400 MHz, CD2Cl2, 203 K): 40 ± 2 (−10.64 ppm). Synthesis of [(Os(H···H){xant(PiPr2)2})2(μ-Cl)2][BF4]2 (3). HBF4· OEt2 (37 μL; 0.272 mmol) was added to a solution of [OsH3Cl{xant(PiPr2)2}] (1; 130 mg; 0,194 mmol) in dichloromethane (5 mL). After the mixture was stirred for 1 h, the solvent was evaporated and the residue was treated with cold pentane (2 × 1 mL) to afford a pale yellow solid which was dried overnight under reduced pressure. Yield: 113 mg (95%). Yellow crystals suitable for X-ray diffraction were obtained by vapor diffusion of pentane into a CH2Cl2 solution. Anal. Calcd for C54H84B2Cl2F8O2Os2P4: C, 42.84; H, 5.59. Found: C, 42.60; H, 5.82. IR (ATR, cm−1): ν(BF) 1050 (s). 1H NMR (500 MHz, CD2Cl2, 293 K): δ 7.83 (m, 4H, CH-arom), 7.55 (m, 8H, CH-arom), 2.58 (m, 4H, PCH(CH3)2), 2.11 (s, 3H, C(CH3)2), 2.08 (s, 3H, C(CH3)2), 1.99 (m, 4H, PCH(CH3)2), 1.24 (dd, 3JHP = 17.5, 3JHH = 7.0, 12H, PCH(CH3)2), 1.12 (dd, 3JHP = 17.0, 3JHH = 7.0, 12H, PCH(CH3)2), 1.02 (dd, 3JHP = 17.0, 3JHH = 7.3, 12H, PCH(CH3)2), 0.85 (dd, 3JHP = 15.1, 3JHH = 6.9, 12H, PCH(CH3)2), −9.53 (t, 2JHP = 14.5, 4H, OsH). 13C{1H}-APT NMR (75.47 MHz, CD2Cl2, 293 K): δ 159.7 (d, 2JCP = 8.7, C-arom), 137.8 (d, 3JCP = 4.6, C-arom), 130.9, 130.1 (both s, CH-arom), 128.3 (d, 2JCP = 6.1, CH-arom), 120.1 (d, 1 JCP = 40.7, C-arom), 38.5 (s, C(CH3)2), 32.2 (s, C(CH3)2), 32.0 (d, 1 JCP = 33.8, PCH(CH3)2), 25.0 (d, 1JCP = 36.3, PCH(CH3)2), 23.3 (s, C(CH3)2), 22.1 (s, PCH(CH3)2), 21.4 (s, PCH(CH3)2), 19.0 (s, E

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry showed a triplet (JHD = 20 Hz) at −9.87 ppm, corresponding to the deuterated isotopomer. Synthesis of [K(18-crown-6)][OsH5{xant(PiPr2)2}] (8). To a THF (0.6 mL) solution of [OsH4{xant(PiPr2)2}] (7; 40 mg, 0.063 mmol) were added KH (8 mg, 0.200 mmol) and 18-crown-6 (50 mg, 0.190 mmol). The reaction mixture was heated to 70 °C for 14 h and then filtered while hot. The filtrate was cooled to room temperature to afford yellow crystals, which were isolated after removal of the supernatant by decantation and dried overnight under reduced pressure. Yield: 39 mg (66%). Complex 8 can be recrystallized by slow vapor diffusion of pentane into a solution of the compound in THF. Anal. Calcd for C39H69KO7OsP2: C, 49.77; H, 7.39. Found: C, 50.17; H, 7.65. HRMS (electrospray, m/z): calcd for C27H44OOsP2 [M − H]+, 638.2478; found, 638.2479. 1H NMR (400.13 MHz, THFd8, 293 K): 7.19 (d, 3JHH = 7.5, 2H, CH-arom), 7.13 (m, 2H, CHarom) 6.89 (dd, 3JHH = 7.5, 3JHH = 7.5, 2H, CH-arom), 3.51 (s, 24H, CH2-18-crown-6), 2.18 (m, 4H, PCH(CH3)2), 1.53 (s, 6H, C(CH3)2), 1.16 (dvt, 3JHH = 7.0, N = 13.0, 12H, PCH(CH3)2), 1.13 (dvt, N = 11.9, 3JHH = 6.5, 12H, PCH(CH3)2), −12.46 (br, 5H, OsH). 13C{1H} NMR-APT (100.62 MHz, THF-d8, 343 K): δ 158.0 (vt, N = 8.0, Carom), 135.0 (s, C-arom), 132.2 (m, C-arom), 127.6 (s, CH-arom), 121.7 (s, CH-arom), 120.0 (s, CH-arom), 70.5 (s, CH2-18-crown-6), 36.0 (s, C(CH3)2), 29.2 (d, 1JCP = 27.1, PCH(CH3)2), 26.6 (s, C(CH3)2), 20.9 (vt, N = 4.6, PCH(CH3)2), 19.1 (s, PCH(CH3)2). 31 1 P{ H} NMR (161.98 MHz, THF-d8, 293 K): δ 20.6 (s, POP). Reaction of [(Os(H···H){xant(PiPr2)2})2(μ-Cl)2][BF4]2 (3) with KH. [(Os(H···H){xant(PiPr2)2})2(μ-Cl)2][BF4]2 (3; 100 mg, 0.066 mmol) and KH (6 mg, 0.15 mmol) were placed in a Schlenk tube, and THF (5 mL) was added to form a suspension. The reaction mixture was stirred and monitored by NMR, showing complete conversion into [OsH3Cl{xant(PiPr2)2}] (1) after 1 h. Computational Details and Cartesian Coordinates of 2, 5, and 6. All calculations were performed at the DFT level using the B3LYP functional 29 supplemented with Grimme’s dispersion correction D330 as implemented in Gaussian09.23 The Os atom was described by means of an effective core potential SDD for the inner electron31 and its associated double-ζ basis set for the outer electrons, complemented with a set of f-polarization functions.32 The 6-31G** basis set was used for the H, C, O, N, Cl, and P atoms.33 All minima were verified to have no negative frequencies. All geometries were fully optimized in vacuo. Structural Analysis of Complexes 3, 5, 6, and 8. X-ray data were collected on Bruker Smart APEX CCD (5, 6, and 8) and Smart APEX CCD DUO (3) diffractometers using graphite-monochromated Mo Kα radiation (λ = 0.71073 Å). Data were collected over the complete sphere and were corrected for absorption by using a multiscan method applied with the SADABS program. The structures were solved by Patterson or direct methods and refined by full-matrix least squares on F2 with SHELXL97, including isotropic and subsequently anisotropic displacement parameters. The hydrogen atoms were observed in the least-squares Fourier maps or calculated and refined freely or using a restricted riding model. However, the hydride ligands were observed in the difference Fourier maps but did not refine properly; therefore, the osmium− hydride distance was fixed in the refinement (1.59 Å, CCDC). Complex 8 is twined by pseudomerohedry. The structure simulates orthorhombic symmetry but was properly refined in the monoclinic symmetry system with a β value of approximately 90° and refined with the twin law 100, 0−10, 00−1 plus one BASF parameter of ∼0.5. Due to the difficulty of the refinement, restrictions in the displacement parameters were used. In the four structures there are anions and crystallization molecules observed to be disordered, and these were refined with different moieties, restrained geometry, and isotropic thermal parameters. Crystal data for 3: C54H84Cl2O2Os2P4·2BF4·CH2Cl2, mol wt 1598.94, irregular block, yellow (0.19 × 0.12 × 0.08), monoclinic, space group P21/c, a = 12.9473(15) Å, b = 20.565(2) Å, c = 12.0926(14) Å, β = 93.005(2)°, V = 3215.4(6) Å3, Z = 2, Z′ = 0.5, Dcalc = 1.651 g cm−3, F(000) = 1588, T = 100(2) K, μ = 4.275 mm−1; 41253 measured reflections (2θ = 3−58°, ω scans 0.3°), 8442 unique

(Rint = 0.0411); minimum/maximum transmission factors 0.633/ 0.862. Final agreement factors were R1 = 0.0394 (6950 observed reflections, I > 2σ(I)) and wR2 = 0.1061; data/restraints/parameters 8442/35/379; GOF = 0.949. Largest peak and hole: 2.66 (close to osmium atoms) and −1.705 e/ Å3. Crystal data for 5: C24H38ClOOsP2·CF3O3S·CHF3O3S, mol wt 929.28, irregular block, colorless (0.19 × 0.14 × 0.08), monoclinic, space group Cc, a = 16.2471(10) Å, b = 21.4146(13) Å, c = 10.2694(6) Å, β = 104.6620(10)°, V = 3456.6(4) Å3, Z = 4, Z′ = 1. Dcalc = 1.786 g cm−3, F(000) = 1840, T = 100(2) K, μ = 4.053 mm−1; 19163 measured reflections (2θ = 3−58°, ω scans 0.3°), 7949 unique (Rint = 0.0320); minimum/maximum transmission factors 0.687/0.862. Final agreement factors were R1 = 0.0277 (7384 observed reflections, I > 2σ(I)) and wR2 = 0.0555; data/restraints/parameters 7949/6/428; GOF = 1.022. Largest peak and hole: 0.827 and −0.846 e/ Å3. Crystal data for 6: C29H45ClNOOsP2·BF4·C4H10O, mol wt 872.18, irregular block, colorless (0.15 × 0.10 × 0.07), monoclinic, space group P21/m, a = 11.843(3) Å, b = 22.139(5) Å, c = 14.179(3) Å, β = 89.988(3)°, V = 3717.4(14) Å3, Z = 4, Z′ = 1, Dcalc = 1.558 g cm−3, F(000) = 1760, T = 100(2) K, μ = 3.638 mm−1; 32991 measured reflections (2θ = 3−58°, ω scans 0.3°), 7908 unique (Rint = 0.0876); minimum/maximum transmission factors 0.590/0.862. Final agreement factors were R1 = 0.0521 (5862 observed reflections, I > 2σ(I)) and wR2 = 0.1289; data/restraints/parameters 7908/256/417; GOF = 1.135. Largest peak and hole: 1.785 (close to osmium atoms) and −1.406 e/ Å3. Crystal data for 8: C27H45OOsP2·C12H24KO6·0.5C4H8O·0.5C5H12, mol wt 1013.31, irregular block, yellow (0.25 × 0.06 × 0.06), monoclinic, space group P21/n, a = 17.0875(9) Å, b = 15.2536(8) Å, c = 19.5724(10) Å, β = 109.6270(10)°, V = 4805.1(4) Å3, Z = 4, Z′ = 1, Dcalc = 1.401 g cm−3, F(000) = 2100, T = 100(2) K, μ = 2.852 mm−1; 82792 measured reflections (2θ = 3−58°, ω scans 0.3°), 11705 unique (Rint = 0.0583); minimum/maximum transmission factors 0.679/ 0.862. Final agreement factors were R1 = 0.0386 (9412 observed reflections, I > 2σ(I)) and wR2 = 0.0946; data/restraints/parameters 11705/17/510; GOF = 1.032. Largest peak and hole: 2.348 (close to osmium atoms) and −1.381 e/Å3.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.6b02837. Computational details and Cartesian coordinates of 2, 5, and 6 (PDF) Crystallographic details for 3, 5, 6, and 8 (CIF) Theoretical complex coordinates (XYZ)



AUTHOR INFORMATION

Corresponding Author

*E-mail for M.A.E.: [email protected]. ORCID

Miguel A. Esteruelas: 0000-0002-4829-7590 Cristina García-Yebra: 0000-0002-5545-5112 Jaime Martín: 0000-0003-0909-3509 Enrique Oñate: 0000-0003-2094-719X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS Financial support from the MINECO of Spain (Projects CTQ2014-52799-P and CTQ2014-51912-REDC), Gobierno de Aragón (E35), FEDER, and the European Social Fund is acknowledged. F

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry



Rhodium versus Iridium. Inorg. Chem. 2013, 52, 5339−5349. (e) Esteruelas, M. A.; Oliván, M.; Vélez, A. POP-Pincer Silyl Complexes of Group 9: Rhodium versus Iridium. Inorg. Chem. 2013, 52, 12108−12119. (f) Esteruelas, M. A.; Oliván, M.; Vélez, A. Conclusive Evidence on the Mechanism of the Rhodium-Mediated Decyanative Borylation. J. Am. Chem. Soc. 2015, 137, 12321−12329. (g) Esteruelas, M. A.; Nolis, P.; Oliván, M.; Oñate, E.; Vallribera, A.; Vélez, A. Ammonia Borane Dehydrogenation Promoted by a PincerSquare-Planar Rhodium(I) Monohydride: A Stepwise Hydrogen Transfer from the Substrate to the Catalyst. Inorg. Chem. 2016, 55, 7176−7181. (6) (a) Pontiggia, A. J.; Chaplin, A. B.; Weller, A. S. Cationic Iridium Complexes of the Xantphos Ligand. Flexible Coordination Modes and the Isolation of the Hydride Insertion Product with an Alkene. J. Organomet. Chem. 2011, 696, 2870−2876. (b) Dallanegra, R.; Chaplin, A. B.; Weller, A. S. Rhodium Cyclopentyl Phosphine Complexes of Wide-Bite-Angle Ligands DPEphos and Xantphos. Organometallics 2012, 31, 2720−2728. (7) Pawley, R. J.; Moxham, G. L.; Dallanegra, R.; Chaplin, A. B.; Brayshaw, S. K.; Weller, A. S.; Willis, M. C. Controlling Selectivity in Intermolecular Alkene or Aldehyde Hydroacylation Reactions Catalyzed by {Rh(L2)}+ Fragments. Organometallics 2010, 29, 1717−1728. (8) Ren, P.; Pike, S. D.; Pernik, I.; Weller, A. S.; Willis, M. C. Rh-POP Pincer Xantphos Complexes for C-S and C-H Activation. Implications for Carbothiolation Catalysis. Organometallics 2015, 34, 711−723. (9) Johnson, H. C.; Leitao, E. M.; Whitten, G. R.; Manners, I.; LloydJones, G. C.; Weller, A. S. Mechanistic Studies of the Dehydrocoupling and Dehydropolymerization of Amine-Boranes Using a [Rh(Xantphos)]+ Catalyst. J. Am. Chem. Soc. 2014, 136, 9078−9093. (10) Esteruelas, M. A.; López, A. M.; Oliván, M. Polyhydrides of Platinum Group Metals: Nonclassical Interactions and σ-Bond Activation Reactions. Chem. Rev. 2016, 116, 8770−8847. (11) Crabtree, R. H. Dihydrogen Complexation. Chem. Rev. 2016, 116, 8750−8769. (12) (a) Rossin, A.; Peruzzini, M. Ammonia-Borane and AmineBorane Dehydrogenation Mediated by Complex Metal Hydrides. Chem. Rev. 2016, 116, 8848−8872. (b) Bhunya, S.; Malakar, T.; Ganguly, G.; Paul, A. Combining Protons and Hydrides by Homogeneous Catalysis for Controlling the Release of Hydrogen from Ammonia−Borane: Present Status and Challenges. ACS Catal. 2016, 6, 7907−7934. (13) (a) Esteruelas, M. A.; Fernández, I.; López, A. M.; Mora, M.; Oñate, E. Osmium-Promoted Dehydrogenation of Amine-Boranes and B-H Bond Activation of the Resulting Amino-Boranes. Organometallics 2014, 33, 1104−1107. (b) Esteruelas, M. A.; López, A. M.; Mora, M.; Oñ ate, E. Ammonia-Borane Dehydrogenation Promoted by an Osmium Dihydride Complex: Kinetics and Mechanism. ACS Catal. 2015, 5, 187−191. (14) (a) Casarrubios, L.; Esteruelas, M. A.; Larramona, C.; Muntaner, J. G.; Oliván, M.; Oñate, E.; Sierra, M. A. Chelated Assisted MetalMediated N-H Bond Activation of β-Lactams: Preparation of Irida-, Rhoda-, Osma-, and Ruthenatrinems. Organometallics 2014, 33, 1820− 1833. (b) Casarrubios, L.; Esteruelas, M. A.; Larramona, C.; Lledós, A.; Muntaner, J. G.; Oñate, E.; Ortuño, M. A.; Sierra, M. A. Mechanistic Insight into the Facilitation of β-Lactam Fragmentation through Metal Assistance. Chem. - Eur. J. 2015, 21, 16781−16785. (c) Casarrubios, L.; Esteruelas, M. A.; Larramona, C.; Muntaner, J. G.; Oñate, E.; Sierra, M. A. 2-Azetidinones as Precursors of Pincer Ligands: Preparation, Structure, and Spectroscopic Properties of CC′N-Osmium Complexes. Inorg. Chem. 2015, 54, 10998−11006. (15) (a) Esteruelas, M. A.; García-Raboso, J.; Oliván, M.; Oñate, E. N-H and N-C Bond Activation of Pyrimidinic Nucleobases and Nucleosides Promoted by an Osmium Polyhydride. Inorg. Chem. 2012, 51, 5975−5984. (b) Esteruelas, M. A.; García-Raboso, J.; Oliván, M. Reactions of an Osmium-Hexahydride Complex with Cytosine, Deoxycytidine, and Cytidine: The Importance of the Minor Tautomers. Inorg. Chem. 2012, 51, 9522−9528.

REFERENCES

(1) (a) Noyori, R.; Ohkuma, T. Asymmetric Catalysis by Architectural and Functional Molecular Engineering: Practical Chemo- and Stereoselective Hydrogenation of Ketones. Angew. Chem., Int. Ed. 2001, 40, 40−73. (b) Clapham, S. E.; Hadzovic, A.; Morris, R. H. Mechanisms of the H2-Hydrogenation and Transfer Hydrogenation of Polar Bonds Catalyzed by Ruthenium Hydride Complexes. Coord. Chem. Rev. 2004, 248, 2201−2237. (c) Conley, B. L.; Pennington-Boggio, M. K.; Boz, E.; Williams, T. J. Discovery, Applications, and Catalytic Mechanisms of Shvo’s Catalyst. Chem. Rev. 2010, 110, 2294−2312. (d) Gunanathan, C.; Milstein, D. MetalLigand Cooperation by Aromatization-Dearomatization: A New Paradigm in Bond Activation and ″Green″ Catalysis. Acc. Chem. Res. 2011, 44, 588−602. (e) Bajo, S.; Esteruelas, M. A.; López, A. M.; Oñate, E. Osmium-Acyl Decarbonylation Promoted by Tp-Mediated Allenylidene Abstraction: A New Role of the Tp Ligand. Organometallics 2014, 33, 4057−4066. (f) Bolaño, T.; Esteruelas, M. A.; Gay, M. P.; Oñate, E.; Pastor, I. M.; Yus, M. An Acyl-NHC Osmium Cooperative System: Coordination of Small Molecules and Heterolytic B-H and O-H Bond Activation. Organometallics 2015, 34, 3902−3908. (2) (a) Periana, R. A.; Taube, D. J.; Gamble, S.; Taube, H.; Satoh, T.; Fujii, H. Platinum Catalysts for the High-Yield Oxidation of Methane to a Methanol Derivative. Science 1998, 280, 560−564. (b) Hashiguchi, B. G.; Young, K. J. H.; Yousufuddin, M.; Goddard, W. A.; Periana, R. A. Acceleration of Nucleophilic CH Activation by Strongly Basic Solvents. J. Am. Chem. Soc. 2010, 132, 12542−12545. (c) Crabtree, R. H. Creating Ligands with Multiple Personalities. Science 2010, 330, 455−456. (d) Kaim, W. Manifestations of Noninnocent Ligand Behavior. Inorg. Chem. 2011, 50, 9752−9765. (3) (a) Braunstein, P.; Naud, F. Hemilability of Hybrid Ligands and the Coordination Chemistry of Oxazoline-Based Systems. Angew. Chem., Int. Ed. 2001, 40, 680−699. (b) Angell, S. E.; Rogers, C. W.; Zhang, Y.; Wolf, M. O.; Jones, W. E. Hemilabile Coordination Complexes for Sensing Applications. Coord. Chem. Rev. 2006, 250, 1829−1841. (c) Braunstein, P. Bonding and Organic and Inorganic Reactivity of Metal-Coordinated Phosphinoenolates and Related Functional Phosphine-Derived Anions. Chem. Rev. 2006, 106, 134− 159. (d) Weng, Z. Q.; Teo, S. H.; Hor, T. S. A. Metal Unsaturation and Ligand Hemilability in Suzuki Coupling. Acc. Chem. Res. 2007, 40, 676−684. (e) Zhang, W. H.; Chien, S. W.; Hor, T. S. A. Recent Advances in Metal Catalysts with Hybrid Ligands. Coord. Chem. Rev. 2011, 255, 1991−2024. (f) Annibale, V. T.; Song, D. T. Multidentate Actor Ligands as Versatile Platforms for Small Molecule Activation and Catalysis. RSC Adv. 2013, 3, 11432−11449. (4) (a) van der Boom, M. E.; Milstein, D. Cyclometalated PhosphineBased Pincer Complexes: Mechanistic Insight in Catalysis, Coordination, and Bond Activation. Chem. Rev. 2003, 103, 1759−1792. (b) Choi, J.; MacArthur, A. H. R.; Brookhart, M.; Goldman, A. S. Dehydrogenation and Related Reactions Catalyzed by Iridium Pincer Complexes. Chem. Rev. 2011, 111, 1761−1779. (c) Haibach, M. C.; Kundu, S.; Brookhart, M.; Goldman, A. S. Alkane Metathesis by Tandem Alkane-Dehydrogenation-Olefin-Metathesis Catalysis and Related Chemistry. Acc. Chem. Res. 2012, 45, 947−958. (d) Kumar, A.; Zhou, T.; Emge, T. J.; Mironov, O.; Saxton, R. J.; Krogh-Jespersen, K.; Goldman, A. S. Dehydrogenation of n-Alkanes by Solid-Phase Molecular Pincer-Iridium Catalysts. High Yields of α-Olefin Product. J. Am. Chem. Soc. 2015, 137, 9894−9911. (5) (a) Boone, M. P.; Brown, C. C.; Ancelet, T. A.; Stephan, D. W. Interconversion of Ruthenium-O(CH2CH2PCy2)2 Alkylidene and Alkylidyne Hydride Complexes. Organometallics 2010, 29, 4369− 4374. (b) Lumbroso, A.; Koschker, P.; Vautravers, N. R.; Breit, B. Redox-Neutral Atom-Economic Rhodium-Catalyzed Coupling of Terminal Alkynes with Carboxylic Acids Toward Branched Allylic Esters. J. Am. Chem. Soc. 2011, 133, 2386−2389. (c) Pawley, R. J.; Huertos, M. A.; Lloyd-Jones, G. C.; Weller, A. S.; Willis, M. C. Intermolecular Alkyne Hydroacylation. Mechanistic Insight from the Isolation of the Vinyl Intermediate that Precedes Reductive Elimination. Organometallics 2012, 31, 5650−5659. (d) Esteruelas, M. A.; Oliván, M.; Vélez, A. Xantphos-Type Complexes of Group 9: G

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX

Article

Inorganic Chemistry (16) (a) Esteruelas, M. A.; Herrero, J.; López, A. M.; Oliván, M. Alkyne-Coupling Reactions Catalyzed by OsHCl(CO)(PiPr3) in the Presence of Diethylamine. Organometallics 2001, 20, 3202−3205. (b) Barrio, P.; Esteruelas, M. A.; Oñate, E. Reactions of a HexahydrideOsmium Complex with Aldehydes: Double C-Hα ActivationDecarbonylation and Single C-Hα Activation-Hydroxylation Tandem Processes and Catalytic Tishchenko Reactions. Organometallics 2004, 23, 1340−1348. (c) Esteruelas, M. A.; Masamunt, A. B.; Oliván, M.; Oñate, E.; Valencia, M. Aromatic Diosmatricyclic Nitrogen-Containing Compounds. J. Am. Chem. Soc. 2008, 130, 11612−11613. (d) Esteruelas, M. A.; Honczek, N.; Oliván, M.; Oñate, E.; Valencia, M. Direct Access to POP-Type Osmium(II) and Osmium(IV) Complexes: Osmium a Promising Alternative to Ruthenium for the Synthesis of Imines from Alcohols and Amines. Organometallics 2011, 30, 2468− 2471. (17) (a) Crespo, O.; Eguillor, B.; Esteruelas, M. A.; Fernández, I.; García-Raboso, J.; Gómez-Gallego, M.; Martín-Ortiz, M.; Oliván, M.; Sierra, M. A. Synthesis and Characterisation of [6]-Azaosmahelicenes: the First d4-Heterometallahelicenes. Chem. Commun. 2012, 48, 5328− 5330. (b) Alabau, R. G.; Eguillor, B.; Esler, J.; Esteruelas, M. A.; Oliván, M.; Oñate, E.; Tsai, J. Y.; Xia, C. J. CCC-Pincer-NHC Osmium Complexes: New Types of Blue-Green Emissive Neutral Compounds for Organic Light-Emitting Devices (OLEDs). Organometallics 2014, 33, 5582−5596. (c) Eguillor, B.; Esteruelas, M. A.; Fernández, I.; Gómez-Gallego, M.; Lledós, A.; Martín-Ortiz, M.; Oliván, M.; Oñate, E.; Sierra, M. A. Azole Assisted C-H Bond Activation Promoted by an Osmium-Polyhydride: Discerning between N and NH. Organometallics 2015, 34, 1898−1910. (18) Bertoli, M.; Chouale, A.; Gusev, D. G.; Lough, A. J.; Major, Q.; Moore, B. PNP Pincer Osmium Polyhydrides for Catalytic Dehydrogenation of Primary Alcohols. Dalton Trans. 2011, 40, 8941−8949. (19) Liu, S. H.; Huang, X.; Lin, Z. Y.; Lau, C. P.; Jia, G. C. Synthesis and Characterization of Dihydrogen(olefin)osmium Complexes with (E)-Ph2P(CH2)2CHCH(CH2)2PPh2. Eur. J. Inorg. Chem. 2002, 2002, 1697−1702. (20) Kuznetsov, V. F.; Gusev, D. G. Chiral Hydride and Dihydrogen Pincer-Type Complexes of Osmium. Organometallics 2007, 26, 5661− 5666. (21) Alós, J.; Bolaño, T.; Esteruelas, M. A.; Oliván, M.; Oñate, E.; Valencia, M. POP-Pincer Osmium-Polyhydrides: Head-to-Head (Z)Dimerization of Terminal Alkynes. Inorg. Chem. 2013, 52, 6199−6213. (22) (a) Kong, V. C. Y.; Foulkes, F. R.; Kirk, D. W.; Hinatsu, J. T. Development of Hydrogen Storage for Fuel Cell Generators. I: Hydrogen Generation Using Hydrolysis Hydrides. Int. J. Hydrogen Energy 1999, 24, 665−675. (b) Kong, V. C. Y.; Kirk, D. W.; Foulkes, F. R.; Hinatsu, J. T. Development of Hydrogen Storage for Fuel Cell Generators II: Utilization of Calcium Hydride and Lithium Hydride. Int. J. Hydrogen Energy 2003, 28, 205−214. (c) Dincer, I. Green Methods for Hydrogen Production. Int. J. Hydrogen Energy 2012, 37, 1954−1971. (d) Dincer, I.; Acar, C. Review and Evaluation of Hydrogen Production Methods for Better Sustainability. Int. J. Hydrogen Energy 2015, 40, 11094−11111. (23) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision B.01; Gaussian, Inc., Wallingford, CT, 2009.

(24) In addition to ref 21, for osmium complexes previously described with this diphosphine see: (a) Asensio, G.; Cuenca, A. B.; Esteruelas, M. A.; Medio-Simón, M.; Oliván, M.; Valencia, M. Osmium(III) Complexes with POP Pincer Ligands: Preparation from Commercially Available OsCl3·3H2O and Their X-ray Structures. Inorg. Chem. 2010, 49, 8665−8667. (b) Alós, J.; Bolaño, T.; Esteruelas, M. A.; Oliván, M.; Oñate, E.; Valencia, M. POP-Pincer Ruthenium Complexes: d6 Counterparts of Osmium d4 Species. Inorg. Chem. 2014, 53, 1195−1209. (c) Alós, J.; Esteruelas, M. A.; Oliván, M.; Oñate, E.; Puylaert, P. C-H Bond Activation Reactions in Ketones and Aldehydes Promoted by POP-Pincer Osmium and Ruthenium Complexes. Organometallics 2015, 34, 4908−4921. (25) dH−H = 5.815(T1(min)/ν)1/6 and dH−H = 1.44 − 0.0168(JH−D). Morris, R. H. Dihydrogen, Dihydride and in Between: NMR and Structural Properties of Iron Group Complexes. Coord. Chem. Rev. 2008, 252, 2381−2394. (26) The structure has two chemically equivalent but crystallographically independent molecules in the asymmetric unit. (27) (a) Gusev, D. G.; Lough, A. J.; Morris, R. H. New Polyhydride Anions and Proton-Hydride Hydrogen Bonding in their Ion Pairs. Xray Crystal Structure Determinations of Q[mer-Os(H)3(CO)(PiPr3)2], Q = [K(18-crown-6)] and Q = [K(1-aza-18-crown-6)]. J. Am. Chem. Soc. 1998, 120, 13138−13147. (b) Abdur-Rashid, K.; Gusev, D. G.; Lough, A. J.; Morris, R. H. Intermolecular Proton-Hydride Bonding in Ion Pairs: Synthesis and Structural Properties of [K(Q)][MH5((PiPr3)2)] (M = Os, Ru; Q = 18-crown-6, 1-aza-18-crown-6, 1,10-diaza-18-crown-6). Organometallics 2000, 19, 834−843. (28) (a) Howard, J. A. K.; Johnson, O.; Koetzle, T. F.; Spencer, J. L. Crystal and Molecular Structure of Bis(diisopropylphenylphosphine)hexahydridoosmium, [OsH6(PC12H19)2]: Single-Crystal Neutron Diffraction Study at 20 K. Inorg. Chem. 1987, 26, 2930−2933. (b) Buil, M. L.; Cardo, J. J. F.; Esteruelas, M. A.; Fernández, I.; Oñate, E. An Entry to Stable Mixed Phosphine-Osmium-NHC Polyhydrides. Inorg. Chem. 2016, 55, 5062−5070. (29) (a) Lee, C. T.; Yang, W. T.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron-Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (b) Becke, A. D. A New Mixing of Hartree-Fock and Local Density-Functional Theories. J. Chem. Phys. 1993, 98, 1372−1377. (c) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. AbInitio Calculation of Vibrational Absorption and Circular-Dichroism Spectra Using Density-Functional Force-Fields. J. Phys. Chem. 1994, 98, 11623−11627. (d) Alós, J.; Bolaño, T.; Esteruelas, M. A.; Oliván, M.; Oñate, E.; Valencia, M. POP-Pincer Osmium-Polyhydrides: Headto-Head (Z)-Dimerization of Terminal Alkynes. Inorg. Chem. 2013, 52, 6199−6213. (30) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A Consistent and Accurate ab initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (31) Andrae, D.; Haussermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Energy-Adjusted Abinitio Pseudopotentials for the 2nd and 3rd Row Transition-Elements. Theoretica Chimica Acta 1990, 77, 123−141. (32) Ehlers, A. W.; Bohme, M.; Dapprich, S.; Gobbi, A.; Hollwarth, A.; Jonas, V.; Kohler, K. F.; Stegmann, R.; Veldkamp, A.; Frenking, G. A Set of F-Polarization Functions for Pseudo-Potential Basis-Sets of the Transition-Metals Sc-Cu, Y-Ag and La-Au. Chem. Phys. Lett. 1993, 208, 111−114. (33) (a) Hehre, W. J.; Ditchfield, R.; Pople, J. A. Self-Consistent Molecular-Orbital Methods. XII. Further Extensions of Gaussian-Type Basis Sets for Use in Molecular-Orbital Studies of Organic-Molecules. J. Chem. Phys. 1972, 56, 2257−2261. (b) Francl, M. M.; Pietro, W. J.; Hehre, W. J.; Binkley, J. S.; Gordon, M. S.; Defrees, D. J.; Pople, J. A. Self-Consistent Molecular-Orbital Methods 0.23. A Polarization-Type Basis Set for 2nd-Row Elements. J. Chem. Phys. 1982, 77, 3654−3665.

H

DOI: 10.1021/acs.inorgchem.6b02837 Inorg. Chem. XXXX, XXX, XXX−XXX